Sie sind auf Seite 1von 15

Big Bang Theory: Evolution of Our Universe

How was our Universe created? How did it come to be the seemingly infinite place we know of today?
And what will become of it, ages from now? These are the questions that have been puzzling philosophers
and scholars since the beginning the time, and led to some pretty wild and interesting theories. Today, the
consensus among scientists, astronomers and cosmologists is that the Universe as we know it was
created in a massive explosion that not only created the majority of matter, but the physical laws that
govern our ever-expanding cosmos. This is known as The Big Bang Theory.

For almost a century, the term has been bandied about by scholars and non-scholars alike. This should
come as no surprise, seeing as how it is the most accepted theory of our origins. But what exactly does it
mean? How was our Universe conceived in a massive explosion, what proof is there of this, and what
does the theory say about the long-term projections for our Universe?

The basics of the Big Bang theory are fairly simple. In short, the Big Bang hypothesis states that all of
the current and past matter in the Universe came into existence at the same time, roughly 13.8 billion
years ago. At this time, all matter was compacted into a very small ball with infinite density and intense
heat called a Singularity. Suddenly, the Singularity began expanding, and the universe as we know it
began.

While this is not the only modern theory of how the Universe came into being – for example, there is
the Steady State Theory or the Oscillating Universe Theory – it is the most widely accepted and popular.
Not only does the model explain the origin of all known matter, the laws of physics, and the large scale
structure of the Universe, it also accounts for the expansion of the Universe and a broad range of other
phenomena.

Timeline of the Big Bang Theory


Working backwards from the current state of the Universe, scientists have theorized that it must have
originated at a single point of infinite density and finite time that began to expand. After the initial
expansion, the theory maintains that Universe cooled sufficiently to allow the formation of subatomic
particles, and later simple atoms. Giant clouds of these primordial elements later coalesced through
gravity to form stars and galaxies.

This all began roughly 13.8 billion years ago, and is thus considered to be the age of the universe.
Through the testing of theoretical principles, experiments involving particle accelerators and high-energy
states, and astronomical studies that have observed the deep universe, scientists have constructed a
timeline of events that began with the Big Bang and has led to the current state of cosmic evolution.

However, the earliest times of the Universe – lasting from approximately 10-43 to 10-11 seconds after the
Big Bang – are the subject of extensive speculation. Given that the laws of physics as we know them could
not have existed at this time, it is difficult to fathom how the Universe could have been governed. What’s
more, experiments that can create the kinds of energies involved have not yet been conducted. Still, many
theories prevail as to what took place in this initial instant in time, many of which are compatible.

Singularity Epoch
Also known as the Planck Epoch (or Planck Era), this was the earliest known period of the Universe. At
this time, all matter was condensed on a single point of infinite density and extreme heat. During this
period, it is believed that the quantum effects of gravity dominated physical interactions and that no other
physical forces were of equal strength to gravitation.

This Planck period of time extends from point 0 to approximately 10-43seconds, and is so named
because it can only be measured in Planck time. Due to the extreme heat and density of matter, the state
of the universe was highly unstable. It thus began to expand and cool, leading to the manifestation of the
fundamental forces of physics.

From approximately 10-43 second and 10-36, the universe began to cross transition temperatures. It is
here that the fundamental forces that govern the Universe are believed to have begun separating from
each other. The first step in this was the force of gravitation separating from gauge forces, which account
for strong and weak nuclear forces and electromagnetism.

Then, from 10-36 to 10-32 seconds after the Big Bang, the temperature of the universe was low enough
(1028 K) that the forces of electromagnetism (strong force) and weak nuclear forces (weak interaction)
were able to separate as well, forming two distinct forces.

Inflation Epoch

With the creation of the first fundamental forces of the universe, the Inflation Epoch began, lasting from
10-32 seconds in Planck time to an unknown point. Most cosmological models suggest that the Universe at
this point was filled homogeneously with a high-energy density, and that the incredibly high temperatures
and pressure gave rise to rapid expansion and cooling.

This began at 10-37 seconds, where the phase transition that caused for the separation of forces also
led to a period where the universe grew exponentially. It was also at this point in time that baryogenesis
occurred, which refers to a hypothetical event where temperatures were so high that the random motions
of particles occurred at relativistic speeds.

As a result of this, particle–antiparticle pairs of all kinds were being continuously created and destroyed
in collisions, which is believed to have led to the predominance of matter over antimatter in the present
universe. After inflation stopped, the universe consisted of a quark–gluon plasma, as well as all other
elementary particles. From this point onward, the Universe began to cool and matter coalesced and
formed.

Cooling Epoch
As the universe continued to decrease in density and temperature, the energy of each particle began to
decrease and phase transitions continued until the fundamental forces of physics and elementary particles
changed into their present form. Since particle energies would have dropped to values that can be
obtained by particle physics experiments, this period onward is subject to less speculation.

For example, scientists believe that about 10-11 seconds after the Big Bang, particle energies dropped
considerably. At about 10-6 seconds, quarks and gluons combined to form baryons such as protons and
neutrons, and a small excess of quarks over antiquarks led to a small excess of baryons over antibaryons.

Since temperatures were not high enough to create new proton-antiproton pairs (or neutron-anitneutron
pairs), mass annihilation immediately followed, leaving just one in 1010 of the original protons and neutrons
and none of their antiparticles. A similar process happened at about 1 second after the Big Bang for
electrons and positrons. After these annihilations, the remaining protons, neutrons and electrons were no
longer moving relativistically and the energy density of the universe was dominated by photons – and to a
lesser extent, neutrinos.
A few minutes into the expansion, the period known as Big Bang nucleosynthesis also began. Thanks
to temperatures dropping to 1 billion kelvin and the energy densities dropping to about the equivalent of
air, neutrons and protons began to combine to form the universe’s first deuterium (a stable isotope of
Hydrogen) and helium atoms. However, most of the Universe’s protons remained uncombined as
hydrogen nuclei.

After about 379,000 years, electrons combined with these nuclei to form atoms (again, mostly
hydrogen), while the radiation decoupled from matter and continued to expand through space, largely
unimpeded. This radiation is now known to be what constitutes the Cosmic Microwave Background (CMB),
which today is the oldest light in the Universe.

As the CMB expanded, it gradually lost density and energy, and is currently estimated to have a
temperature of 2.7260 ± 0.0013 K(-270.424 °C/ -454.763 °F ) and an energy density
of 0.25 eV/cm3 (or 4.005×10-14 J/m3; 400–500 photons/cm3). The CMB can be seen in all directions at a
distance of roughly 13.8 billion light years, but estimates of its actual distance place it at about 46 billion
light years from the center of the Universe.

Structure Epoch
Over the course of the several billion years that followed, the slightly denser regions of the almost
uniformly distributed matter of the Universe began to become gravitationally attracted to each other. They
therefore grew even denser, forming gas clouds, stars, galaxies, and the other astronomical structures that
we regularly observe today.

This is what is known as the Structure Epoch, since it was during this time that the modern Universe
began to take shape. This consists of visible matter distributed in structures of various sizes, ranging from
stars and planets to galaxies, galaxy clusters, and super clusters – where matter is concentrated – that are
separated by enormous gulfs containing few galaxies.

The details of this process depend on the amount and type of matter in the universe, with cold dark
matter, warm dark matter, hot dark matter, and baryonic matter being the four suggested types. However,
the Lambda-Cold Dark Matter model (Lambda-CDM), in which the dark matter particles moved slowly
compared to the speed of light, is the considered to be the standard model of Big Bang cosmology, as it
best fits the available data.

In this model, cold dark matter is estimated to make up about 23% of the matter/energy of the universe,
while baryonic matter makes up about 4.6%. The Lambda refers to the Cosmological Constant, a theory
originally proposed by Albert Einstein that attempted to show that the balance of mass-energy in the
universe was static. In this case, it is associated with Dark Energy, which served to accelerate the
expansion of the universe and keep its large-scale structure largely uniform.
Diagram showing the Lambda-CBR universe, from the Big Bang to the the current era. Credit: Alex
Mittelmann/Coldcreation

Long-term Predictions for the Future of the Universe


Hypothesizing that the Universe had a starting point naturally gives rise to questions about a possible
end point. If the Universe began as a tiny point of infinite density that started to expand, does that mean it
will continue to expand indefinitely? Or will it one day run out of expansive force, and begin retreating
inward until all matter crunches back into a tiny ball?

Answering this question has been a major focus of cosmologists ever since the debate about which
model of the Universe was the correct one began. With the acceptance of the Big Bang Theory, but prior
to the observation of Dark Energy in the 1990s, cosmologists had come to agree on two scenarios as
being the most likely outcomes for our Universe.

In the first, commonly known as the “Big Crunch” scenario, the universe will reach a maximum size and
then begin to collapse in on itself. This will only be possible if the mass density of the Universe is greater
than the critical density. In other words, as long as the density of matter remains at or above a certain
value (1-3 ×10-26 kg of matter per m³), the Universe will eventually contract.

Alternatively, if the density in the universe were equal to or below the critical density, the expansion
would slow down but never stop. In this scenario, known as the “Big Freeze”, the Universe would go on
until star formation eventually ceased with the consumption of all the interstellar gas in each galaxy.
Meanwhile, all existing stars would burn out and become white dwarfs, neutron stars, and black holes.

Very gradually, collisions between these black holes would result in mass accumulating into larger and
larger black holes. The average temperature of the universe would approach absolute zero, and black
holes would evaporate after emitting the last of their Hawking radiation. Finally, the entropy of the universe
would increase to the point where no organized form of energy could be extracted from it (a scenarios
known as “heat death”).

Modern observations, which include the existence of Dark Energy and its influence on cosmic
expansion, have led to the conclusion that more and more of the currently visible universe will pass
beyond our event horizon (i.e. the CMB, the edge of what we can see) and become invisible to us. The
eventual result of this is not currently known, but “heat death” is considered a likely end point in this
scenario too.

Other explanations of dark energy, called phantom energy theories, suggest that ultimately galaxy
clusters, stars, planets, atoms, nuclei, and matter itself will be torn apart by the ever-increasing expansion.
This scenario is known as the “Big Rip”, in which the expansion of the Universe itself will eventually be its
undoing.

History of the Big Bang Theory


The earliest indications of the Big Bang occurred as a result of deep-space observations conducted in
the early 20th century. In 1912, American astronomer Vesto Slipher conducted a series of observations of
spiral galaxies (which were believed to be nebulae) and measured their Doppler Redshift. In almost all
cases, the spiral galaxies were observed to be moving away from our own.

In 1922, Russian cosmologist Alexander Friedman developed what are known as the Friedman
equations, which were derived from Einstein’s equations for general relativity. Contrary to Einstein’s was
advocating at the time with his a Cosmological Constant, Friedman’s work showed that the universe was
likely in a state of expansion?

In 1924, Edwin Hubble’s measurement of the great distance to the nearest spiral nebula showed that
these systems were indeed other galaxies. At the same time, Hubble began developing a series of
distance indicators using the 100-inch (2.5 m) Hooker telescope at Mount Wilson Observatory. And by
1929, Hubble discovered a correlation between distance and recession velocity – which is now known
as Hubble’s law.

And then in 1927, Georges Lemaitre, a Belgian physicist and Roman Catholic priest, independently
derived the same results as Friedman’s equations and proposed that the inferred recession of the galaxies
was due to the expansion of the universe. In 1931, he took this further, suggesting that the current
expansion of the Universe meant that the father back in time one went, the smaller the Universe would be.
At some point in the past, he argued, the entire mass of the universe would have been concentrated into a
single point from which the very fabric of space and time originated.

These discoveries triggered a debate between physicists throughout the 1920s and 30s, with the
majority advocating that the universe was in a steady state. In this model, new matter is continuously
created as the universe expands, thus preserving the uniformity and density of matter over time. Among
these scientists, the idea of a Big Bang seemed more theological than scientific, and accusations of bias
were made against Lemaitre based on his religious background.

Other theories were advocated during this time as well, such as the Milne Model and the Oscillary
Universe model. Both of these theories were based on Einstein’s theory of general relativity (the latter
being endorsed by Einstein himself), and held that the universe follows infinite, or indefinite, self-sustaining
cycles.

After World War II, the debate came to a head between proponents of the Steady State Model (which
had come to be formalized by astronomer Fred Hoyle) and proponents of the Big Bang Theory – which
was growing in popularity. Ironically, it was Hoyle who coined the phrase “Big Bang” during a BBC Radio
broadcast in March 1949, which was believed by some to be a pejorative dismissal (which Hoyle denied).

Eventually, the observational evidence began to favor Big Bang over Steady State. The discovery and
confirmation of the cosmic microwave background radiation in 1965 secured the Big Bang as the best
theory of the origin and evolution of the universe. From the late 60s to the 1990s, astronomers and
cosmologist made an even better case for the Big Bang by resolving theoretical problems it raised.

These included papers submitted by Stephen Hawking and other physicists that showed that
singularities were an inevitable initial condition of general relativity and a Big Bang model of cosmology. In
1981, physicist Alan Goth theorized of a period of rapid cosmic expansion (aka. the “Inflation” Epoch) that
resolved other theoretical problems.

The 1990s also saw the rise of Dark Energy as an attempt to resolve outstanding issues in cosmology.
In addition to providing an explanation as to the universe’s missing mass (along with Dark Matter,
originally proposed in 1932 by Jan Oort), it also provided an explanation as to why the universe is still
accelerating, as well as offering a resolution to Einstein’s Cosmological Constant.
Significant progress was made thanks to advances in telescopes, satellites, and computer simulations,
which have allowed astronomers and cosmologists to see more of the universe and gain a better
understanding of its true age. The introduction of space telescopes – such as the Cosmic Background
Explorer (COBE), the Hubble Space Telescope, Wilkinson Microwave Anisotropy Probe (WMAP) and
the Planck Observatory – have also been of immeasurable value.

Today, cosmologists have fairly precise and accurate measurements of many of the parameters of the
Big Bang Theory model, not to mention the age of the Universe itself. And it all began with the noted
observation that massive stellar objects, many light years distant, were slowly moving away from us. And
while we still are not sure how it will all end, we do know that on a cosmological scale, that won’t be for a
long, LONG time!

Stellar Formation
Stellar evolution is the process by which a star changes over the course of time. Depending on the mass of
the star, its lifetime can range from a few million years for the most massive to trillions of years for the least
massive, which is considerably longer than the age of the universe. The table shows the lifetimes of stars as
a function of their masses. All stars are born from collapsing clouds of gas and dust, often
called nebulae or molecular clouds. Over the course of millions of years, these proto stars settle down into a
state of equilibrium, becoming what is known as a main-sequence star.
Nuclear fusion powers a star for most of its life. Initially the energy is generated by the fusion of hydrogen
atoms at the core of the main-sequence star. Later, as the preponderance of atoms at the core
becomes helium, stars like the Sun begin to fuse hydrogen along a spherical shell surrounding the core.
This process causes the star to gradually grow in size, passing through the sub giant stage until it reaches
the red giant phase. Stars with at least half the mass of the Sun can also begin to generate energy through
the fusion of helium at their core, whereas more-massive stars can fuse heavier elements along a series of
concentric shells. Once a star like the Sun has exhausted its nuclear fuel, its core collapses into a
dense white dwarf and the outer layers are expelled as a planetary nebula. Stars with around ten or more
times the mass of the Sun can explode in a supernova as their inert iron cores collapse into an extremely
dense neutron star or black hole. Although the universe is not old enough for any of the smallest red
dwarfs to have reached the end of their lives, stellar models suggest they will slowly become brighter and
hotter before running out of hydrogen fuel and becoming low-mass white dwarfs.
Stellar evolution is not studied by observing the life of a single star, as most stellar changes occur too slowly
to be detected, even over many centuries. Instead, astrophysicists come to understand how stars evolve by
observing numerous stars at various points in their lifetime, and by simulating stellar
structure using computer models.

Protostar
Stellar evolution starts with the gravitational collapse of a giant molecular cloud. Typical giant
molecular clouds are roughly 100 light-years (9.5×1014 km) across and contain up to 6,000,000 solar
masses (1.2×1037 kg). As it collapses, a giant molecular cloud breaks into smaller and smaller pieces. In
each of these fragments, the collapsing gas releases gravitational potential energy as heat. As its
temperature and pressure increase, a fragment condenses into a rotating sphere of superhot gas known
as a protostar.
A protostar continues to grow by accretion of gas and dust from the molecular cloud, becoming
a pre-main-sequence star as it reaches its final mass. Further development is determined by its mass.
Mass is typically compared to the mass of the Sun: 1.0 M☉ (2.0×1030 kg) means 1 solar mass.

Protostars are encompassed in dust, and are thus more readily visible at infrared wavelengths.
Observations from the Wide-field Infrared Survey Explorer (WISE) have been especially important for
unveiling numerous Galactic protostars and their parent star clusters.[4][5]

Brown dwarfs and sub-stellar objects


Protostars with masses less than roughly 0.08 M☉ (1.6×1029 kg) never reach temperatures high
enough for nuclear fusion of hydrogen to begin. These are known as brown dwarfs. The International
Astronomical Uniondefines brown dwarfs as stars massive enough to fuse deuterium at some point in their
lives (13 Jupiter masses(MJ), 2.5 × 1028 kg, or 0.0125 M☉). Objects smaller than 13 MJ are classified as
sub-brown dwarfs (but if they orbit around another stellar object they are classified as planets).[6] Both
types, deuterium-burning and not, shine dimly and die away slowly, cooling gradually over hundreds of
millions of years.
A dense starfield in Sagittarius

Main sequence
For a more-massive proto star, the core temperature will eventually reach 10 million kelvin,
initiating the proton–proton chain reaction and allowing hydrogen to fuse, first to deuterium and then
to helium. In stars of slightly over 1 M☉ (2.0×1030 kg), the carbon–nitrogen–oxygen fusion reaction (CNO
cycle) contributes a large portion of the energy generation. The onset of nuclear fusion leads relatively
quickly to a hydrostatic equilibrium in which energy released by the core maintains a high gas pressure,
balancing the weight of the star's matter and preventing further gravitational collapse. The star thus
evolves rapidly to a stable state, beginning the main-sequence phase of its evolution.
A new star will sit at a specific point on the main sequence of the Hertzsprung–Russell diagram, with the
main-sequence spectral type depending upon the mass of the star. Small, relatively cold, low-mass red
dwarfs fuse hydrogen slowly and will remain on the main sequence for hundreds of billions of years or
longer, whereas massive, hot O-type stars will leave the main sequence after just a few million years. A
mid-sized yellow dwarf star, like the Sun, will remain on the main sequence for about 10 billion years. The
Sun is thought to be in the middle of its main sequence lifespan.
The evolutionary tracks of stars with different initial masses on the Hertzsprung–Russell diagram. The tracks start
once the star has evolved to the main sequence and stop when fusion stops (for massive stars) and at the end of
the red giant branch (for stars 1 M☉ and less)
A yellow track is shown for the Sun, which will become a red giant after its main-sequence phase ends before
expanding further along the asymptotic giant branch, which will be the last phase in which the Sun undergoes
fusion.ends before expanding further along the asymptotic giant branch, which will be the last phase in which the Sun
undergoes fusion.

Mature stars
Eventually
the core exhausts its
supply of hydrogen and
the star begins to
evolve off of the main
sequence. Without the

outward pressure generated by the fusion of hydrogen to counteract the force of gravity the core contracts
until either electron degeneracy pressure becomes sufficient to oppose gravity or the core becomes hot
enough (around 100 MK) for helium fusion to begin. Which of these happens first depends upon the star's
mass.

Low-mass stars
What happens after a low-mass star ceases to produce energy through fusion has not been
directly observed; the universe is around 13.8 billion years old, which is less time (by several orders of
magnitude, in some cases) than it takes for fusion to cease in such stars.
Recent astrophysical models suggest that red dwarfs of 0.1 M☉ may stay on the main sequence for some
six to twelve trillion years, gradually increasing in both temperature and luminosity, and take several
hundred billion more to collapse, slowly, into a white dwarf. Such stars will not become red giants as the
whole star is a convection zone and it will not develop a degenerate helium core with a shell burning
hydrogen. Instead, hydrogen fusion will proceed until almost the whole star is helium.

Internal structures of main-sequence stars, convection zones with arrowed cycles and radiative zones with red
flashes. To the left a low-mass red dwarf, in the center a mid-sized yellow dwarf and at the right a massive blue-
white main-sequence star.

Slightly more massive starsdo expand into red giants, but their helium cores are not massive enough to
reach the temperatures required for helium fusion so they never reach the tip of the red giant branch.
When hydrogen shell burning finishes, these stars move directly off the red giant branch like a post-
asymptotic-giant-branch (AGB) star, but at lower luminosity, to become a white dwarf. A star with an initial
mass above about 0.8 M☉ will be able to reach temperatures high enough to fuse helium, and these "mid-
sized" stars go on to further stages of evolution beyond the red giant branch.

Mid-sized stars

The evolutionary track of a solar mass, solar metallicity, star from main sequence to post-AGB

Stars of roughly 0.8–10 M☉ become red giants, which are large non-main-sequence stars
of stellar classification K or M. Red giants lie along the right edge of the Hertzsprung–Russell diagram due
to their red color and large luminosity. Examples include Aldebaran in the
constellation Taurus and Arcturus in the constellation of Bootees.
Mid-sized stars are red giants during two different phases of their post-main-sequence
evolution: red-giant-branch stars, with inert cores made of helium and hydrogen-burning shells, and
asymptotic-giant-branch stars, with inert cores made of carbon and helium-burning shells inside the
hydrogen-burning shells. Between these two phases, stars spend a period on the horizontal branch with a
helium-fusing core. Many of these helium-fusing stars cluster towards the cool end of the horizontal branch
as K-type giants and are referred to as red clump giants.
When a star exhausts the hydrogen in its core, it leaves the main sequence and begins to fuse
hydrogen in a shell outside the core. The core increases in mass as the shell produces more helium.
Depending on the mass of the helium core, this continues for several million to one or two billion years,
with the star expanding and cooling at a similar or slightly lower luminosity to its main sequence state.
Eventually either the core becomes degenerate, in stars around the mass of the sun, or the outer layers
cool sufficiently to become opaque, in more massive stars. Either of these changes cause the hydrogen
shell to increase in temperature and the luminosity of the star to increase, at which point the star expands
onto the red giant branch.
Red-giant-branch

Typical stellar evolution for 0.8-8 M☉

The expanding outer layers of the star are convective, with the material being mixed by
turbulence from near the fusing regions up to the surface of the star. For all but the lowest-mass stars, the
fused material has remained deep in the stellar interior prior to this point, so the convicting envelope
makes fusion products visible at the star's surface for the first time. At this stage of evolution, the results
are subtle, with the largest effects, alterations to the isotopes of hydrogen and helium, being unobservable.
The effects of the CNO cycle appear at the surface during the first dredge-up, with lower 12C/13C ratios and
altered proportions of carbon and nitrogen. These are detectable with spectroscopy and have been
measured for many evolved stars.
The helium core continues to grow on the red giant branch. It is no longer in thermal
equilibrium, either degenerate or above the Schoenberg-Chandrasekhar limit, so it increases in
temperature which causes the rate of fusion in the hydrogen shell to increase. The star increases in
luminosity towards the tip of the red-giant branch. Red giant branch stars with a degenerate helium core all
reach the tip with very similar core masses and very similar luminosities, although the more massive of the
red giants become hot enough to ignite helium fusion before that point.

Horizontal branch
In the helium cores of stars in the 0.8 to 2.0 solar mass range, which are largely supported
by electron degeneracy pressure, helium fusion will ignite on a timescale of days in a helium flash. In the
no degenerate cores of more massive stars, the ignition of helium fusion occurs relatively slowly with no
flash. The nuclear power released during the helium flash is very large, on the order of 108 times the
luminosity of the Sun for a few days. And 1011 times the luminosity of the Sun (roughly the luminosity of
the Milky Way Galaxy) for a few seconds. However, the energy is consumed by the thermal expansion of
the initially degenerate core and thus cannot be seen from outside the star. Due to the expansion of the
core, the hydrogen fusion in the overlying layers slows and total energy generation decreases. The star
contracts, although not all the way to the main sequence, and it migrates to the horizontal branch on the
Hertzsprung–Russell diagram, gradually shrinking in radius and increasing its surface temperature.
Core helium flash stars evolve to the red end of the horizontal branch but do not migrate to
higher temperatures before they gain a degenerate carbon-oxygen core and start helium shell burning.
These stars are often observed as a red clump of stars in the colour-magnitude diagram of a cluster, hotter
and less luminous than the red giants. Higher-mass stars with larger helium cores move along the
horizontal branch to higher temperatures, some becoming unstable pulsating stars in the yellow instability
strip (RR Lyrae variables), whereas some become even hotter and can form a blue tail or blue hook to the
horizontal branch. The morphology of the horizontal branch depends on parameters such as metallicity,
age, and helium content, but the exact details are still being modelled.

Asymptotic-giant-branch
After a star has consumed the helium at the core, hydrogen and helium fusion continues in
shells around a hot core of carbon and oxygen. The star follows the asymptotic giant branch on the
Hertzsprung–Russell diagram, paralleling the original red giant evolution, but with even faster energy
generation (which lasts for a shorter time Although helium is being burnt in a shell, the majority of the
energy is produced by hydrogen burning in a shell further from the core of the star. Helium from these
hydrogen burning shells drops towards the center of the star and periodically the energy output from the
helium shell increases dramatically. This is known as a thermal pulse and they occur towards the end of
the asymptotic-giant-branch phase, sometimes even into the post-asymptotic-giant-branch phase.
Depending on mass and composition, there may be several to hundreds of thermal pulses.
There is a phase on the ascent of the asymptotic-giant-branch where a deep convective zone
forms and can bring carbon from the core to the surface. This is known as the second dredge up, and in
some stars there may even be a third dredge up. In this way a carbon star is formed, very cool and
strongly reddened stars showing strong carbon lines in their spectra. A process known as hot bottom
burning may convert carbon into oxygen and nitrogen before it can be dredged to the surface, and the
interaction between these processes determines the observed luminosities and spectra of carbon stars in
particular clusters.
Another well-known class of asymptotic-giant-branch stars are the Mira variables, which pulsate
with well-defined periods of tens to hundreds of days and large amplitudes up to about 10 magnitudes (in
the visual, total luminosity changes by a much smaller amount). In more-massive stars the stars become
more luminous and the pulsation period is longer, leading to enhanced mass loss, and the stars become
heavily obscured at visual wavelengths. These stars can be observed as OH/IR stars, pulsating in the
infra-red and showing OH maser activity. These stars are clearly oxygen rich, in contrast to the carbon
stars, but both must be produced by dredge ups.

Post-AGB

The Cat's Eye Nebula, a planetary nebula formed by the death of a star with about the same mass as the Sun

These mid-range stars ultimately reach the tip of the asymptotic-giant-branch and run out of fuel
for shell burning. They are not sufficiently massive to start full-scale carbon fusion, so they contract again,
going through a period of post-asymptotic-giant-branch superwind to produce a planetary nebula with an
extremely hot central star. The central star then cools to a white dwarf. The expelled gas is relatively rich
in heavy elements created within the star and may be particularly oxygen or carbon enriched, depending
on the type of the star. The gas builds up in an expanding shell called a circumstellar envelope and cools
as it moves away from the star, allowing dust particles and molecules to form. With the high infrared
energy input from the central star, ideal conditions are formed in these circumstellar envelopes
for maser excitation.
It is possible for thermal pulses to be produced once post-asymptotic-giant-branch evolution
has begun, producing a variety of unusual and poorly understood stars known as born-again asymptotic-
giant-branch stars. These may result in extreme horizontal-branch stars (subdwarf B stars), hydrogen
deficient post-asymptotic-giant-branch stars, variable planetary nebula central stars, and R Coronae
Borealis variables.
Massive stars

Reconstructed image of Antares, a red supergiant

In massive stars, the core is already large enough at the onset of the hydrogen burning shell
that helium ignition will occur before electron degeneracy pressure has a chance to become prevalent.
Thus, when these stars expand and cool, they do not brighten as much as lower-mass stars; however,
they were much brighter than lower-mass stars to begin with, and are thus still brighter than the red giants
formed from less-massive stars. These stars are unlikely to survive as red supergiants; instead they will
destroy themselves as type II supernovas.
Extremely massive stars (more than approximately 40 M☉), which are very luminous and thus
have very rapid stellar winds, lose mass so rapidly due to radiation pressure that they tend to strip off their
own envelopes before they can expand to become red supergiants, and thus retain extremely high surface
temperatures (and blue-white color) from their main-sequence time onwards. The largest stars of the
current generation are about 100-150 M☉because the outer layers would be expelled by the extreme
radiation. Although lower-mass stars normally do not burn off their outer layers so rapidly, they can
likewise avoid becoming red giants or red supergiants if they are in binary systems close enough so that
the companion star strips off the envelope as it expands, or if they rotate rapidly enough so that convection
extends all the way from the core to the surface, resulting in the absence of a separate core and envelope
due to thorough mixing.
The core grows hotter and denser as it gains material from fusion of hydrogen at the base of the
envelope. In all massive stars, electron degeneracy pressure is insufficient to halt collapse by itself, so as
each major element is consumed in the center, progressively heavier elements ignite, temporarily halting
collapse. If the core of the star is not too massive (less than approximately 1.4 M☉, taking into account
mass loss that has occurred by this time), it may then form a white dwarf (possibly surrounded by a
planetary nebula) as described above for less-massive stars, with the difference that the white dwarf is
composed chiefly of oxygen, neon, and magnesium.
The onion-like layers of a massive, evolved star just before core collapse. (Not to scale.)

Above a certain mass (estimated at approximately 2.5 M☉ and whose star's progenitor was
around 10 M☉), the core will reach the temperature (approximately 1.1 gigakelvins) at which neon partially
breaks down to form oxygen and helium, the latter of which immediately fuses with some of the remaining
neon to form magnesium; then oxygen fuses to form sulfur, silicon, and smaller amounts of other
elements. Finally, the temperature gets high enough that any nucleus can be partially broken down, most
commonly releasing an alpha particle (helium nucleus) which immediately fuses with another nucleus, so
that several nuclei are effectively rearranged into a smaller number of heavier nuclei, with net release of
energy because the addition of fragments to nuclei exceeds the energy required to break them off the
parent nuclei.
A star with a core mass too great to form a white dwarf but insufficient to achieve sustained conversion of
neon to oxygen and magnesium, will undergo core collapse (due to electron capture) before achieving
fusion of the heavier elements. Both heating and cooling caused by electron capture onto minor
constituent elements (such as aluminum and sodium) prior to collapse may have a significant impact
on total energy generation within the star shortly before collapse. This may produce a noticeable effect on
the abundance of elements and isotopes ejected in the subsequent supernova.

Supernova

The Crab Nebula, the shattered remnants of a star which exploded as a supernova, the light of which reached Earth
in 1054 AD
Once the nucleo synthesis process arrives at iron-56, the continuation of this process
consumes energy (the addition of fragments to nuclei releases less energy than required to break them off
the parent nuclei). If the mass of the core exceeds the Chandrasekhar limit, electron degeneracy
pressure will be unable to support its weight against the force of gravity, and the core will undergo sudden,
catastrophic collapse to form a neutron star or (in the case of cores that exceed the Tolman-Oppenheimer-
Volkoff limit), a black hole. Through a process that is not completely understood, some of the gravitational
potential energy released by this core collapse is converted into a Type Ib, Type Ic, or Type II supernova.
It is known that the core collapse produces a massive surge of neutrinos, as observed with supernova SN
1987A. The extremely energetic neutrinos fragment some nuclei; some of their energy is consumed in
releasing nucleons, including neutrons, and some of their energy is transformed into heat and kinetic
energy, thus augmenting the shock wave started by rebound of some of the infalling material from the
collapse of the core. Electron capture in very dense parts of the infalling matter may produce additional
neutrons. Because some of the rebounding matter is bombarded by the neutrons, some of its nuclei
capture them, creating a spectrum of heavier-than-iron material including the radioactive elements up to
(and likely beyond) uranium. Although non-exploding red giants can produce significant quantities of
elements heavier than iron using neutrons released in side reactions of earlier nuclear reactions, the
abundance of elements heavier than iron (and in particular, of certain isotopes of elements that have
multiple stable or long-lived isotopes) produced in such reactions is quite different from that produced in a
supernova. Neither abundance alone matches that found in the Solar System, so both supernovae and
ejection of elements from red giants are required to explain the observed abundance of heavy elements
and isotopes thereof.
The energy transferred from collapse of the core to rebounding material not only generates
heavy elements, but provides for their acceleration well beyond escape velocity, thus causing a Type Ib,
Type Ic, or Type II supernova. Current understanding of this energy transfer is still not satisfactory;
although current computer models of Type Ib, Type Ic, and Type II supernovae account for part of the
energy transfer, they are not able to account for enough energy transfer to produce the observed ejection
of material. However, neutrino oscillations may play an important role in the energy transfer problem as
they not only affect the energy available in a particular flavour of neutrinos but also through other general-
relativistic effects on neutrinos.
Some evidence gained from analysis of the mass and orbital parameters of binary neutron stars
(which require two such supernovae) hints that the collapse of an oxygen-neon-magnesium core may
produce a supernova that differs observably (in ways other than size) from a supernova produced by the
collapse of an iron core.[
The most massive stars that exist today may be completely destroyed by a supernova with an
energy greatly exceeding its gravitational binding energy. This rare event, caused by pair-instability, leaves
behind no black hole remnant. In the past history of the universe, some stars were even larger than the
largest that exists today, and they would immediately collapse into a black hole at the end of their lives,
due to photodisintegration.

Stellar evolution of low-mass (left cycle) and high-mass (right cycle) stars, with examples in italics

Stellar remnants
After a star has burned out its fuel supply, its remnants can take one of three forms, depending
on the mass during its lifetime.
White and black dwarfs
For a star of 1 M☉, the resulting white dwarf is of about 0.6 M☉, compressed into approximately the volume
of the Earth. White dwarfs are stable because the inward pull of gravity is balanced by the degeneracy
pressure of the star's electrons, a consequence of the Pauli exclusion principle. Electron degeneracy
pressure provides a rather soft limit against further compression; therefore, for a given chemical
composition, white dwarfs of higher mass have a smaller volume. With no fuel left to burn, the star radiates
its remaining heat into space for billions of years.
A white dwarf is very hot when it first forms, more than 100,000 K at the surface and even hotter
in its interior. It is so hot that a lot of its energy is lost in the form of neutrinos for the first 10 million years of
its existence, but will have lost most of its energy after a billion years.
The chemical composition of the white dwarf depends upon its mass. A star of a few solar
masses will ignite carbon fusion to form magnesium, neon, and smaller amounts of other elements,
resulting in a white dwarf composed chiefly of oxygen, neon, and magnesium, provided that it can lose
enough mass to get below the Chandrasekhar limit (see below), and provided that the ignition of carbon is
not so violent as to blow the star apart in a supernova. A star of mass on the order of magnitude of the
Sun will be unable to ignite carbon fusion, and will produce a white dwarf composed chiefly of carbon and
oxygen, and of mass too low to collapse unless matter is added to it later (see below). A star of less than
about half the mass of the Sun will be unable to ignite helium fusion (as noted earlier), and will produce a
white dwarf composed chiefly of helium.
In the end, all that remains is a cold dark mass sometimes called a black dwarf. However, the
universe is not old enough for any black dwarfs to exist yet.
If the white dwarf's mass increases above the Chandrasekhar limit, which is 1.4 M☉ for a white
dwarf composed chiefly of carbon, oxygen, neon, and/or magnesium, then electron degeneracy pressure
fails due to electron capture and the star collapses. Depending upon the chemical composition and pre-
collapse temperature in the center, this will lead either to collapse into a neutron star or runaway ignition of
carbon and oxygen. Heavier elements favor continued core collapse, because they require a higher
temperature to ignite, because electron capture onto these elements and their fusion products is easier;
higher core temperatures favor runaway nuclear reaction, which halts core collapse and leads to a Type Ia
supernova. These supernovae may be many times brighter than the Type II supernova marking the death
of a massive star, even though the latter has the greater total energy release. This instability to collapse
means that no white dwarf more massive than approximately 1.4 M☉ can exist (with a possible minor
exception for very rapidly spinning white dwarfs, whose centrifugal force due to rotation partially
counteracts the weight of their matter). Mass transfer in a binary system may cause an initially stable white
dwarf to surpass the Chandrasekhar limit.
If a white dwarf forms a close binary system with another star, hydrogen from the larger companion may
accrete around and onto a white dwarf until it gets hot enough to fuse in a runaway reaction at its surface,
although the white dwarf remains below the Chandrasekhar limit. Such an explosion is termed a nova.

Neutron stars

Bubble-like shock wave still expanding from a supernova explosion 15,000 years ago.
Ordinarily, atoms are mostly electron clouds by volume, with very compact nuclei at the center
(proportionally, if atoms were the size of a football stadium, their nuclei would be the size of dust mites).
When a stellar core collapses, the pressure causes electrons and protons to fuse by electron capture.
Without electrons, which keep nuclei apart, the neutrons collapse into a dense ball (in some ways like a
giant atomic nucleus), with a thin overlying layer of degenerate matter (chiefly iron unless matter of
different composition is added later). The neutrons resist further compression by the Pauli Exclusion
Principle, in a way analogous to electron degeneracy pressure, but stronger.
These stars, known as neutron stars, are extremely small—on the order of radius 10 km, no
bigger than the size of a large city—and are phenomenally dense. Their period of rotation shortens
dramatically as the stars shrink (due to conservation of angular momentum); observed rotational periods of
neutron stars range from about 1.5 milliseconds (over 600 revolutions per second) to several
seconds. When these rapidly rotating stars' magnetic poles are aligned with the Earth, we detect a pulse
of radiation each revolution. Such neutron stars are called pulsars, and were the first neutron stars to be
discovered. Though electromagnetic radiation detected from pulsars is most often in the form of radio
waves, pulsars have also been detected at visible, X-ray, and gamma ray wavelengths

Black holes
If the mass of the stellar remnant is high enough, the neutron degeneracy pressure will be
insufficient to prevent collapse below the Schwarzschild radius. The stellar remnant thus becomes a black
hole. The mass at which this occurs is not known with certainty, but is currently estimated at between 2
and 3 M☉.
Black holes are predicted by the theory of general relativity. According to classical general relativity, no
matter or information can flow from the interior of a black hole to an outside observer, although quantum
effects may allow deviations from this strict rule. The existence of black holes in the universe is well
supported, both theoretically and by astronomical observation.
Because the core-collapse mechanism of a supernova is, at present, only partially understood,
it is still not known whether it is possible for a star to collapse directly to a black hole without producing a
visible supernova, or whether some supernovae initially form unstable neutron stars which then collapse
into black holes; the exact relation between the initial mass of the star and the final remnant is also not
completely certain. Resolution of these uncertainties requires the analysis of more supernovae and
supernova remnants.

Models
A stellar evolutionary model is a mathematical model that can be used to compute the evolutionary phases
of a star from its formation until it becomes a remnant. The mass and chemical composition of the star are
used as the inputs, and the luminosity and surface temperature are the only constraints. The model
formulae are based upon the physical understanding of the star, usually under the assumption of
hydrostatic equilibrium. Extensive computer calculations are then run to determine the changing state of
the star over time, yielding a table of data that can be used to determine the evolutionary track of the star
across the Hertzsprung–Russell diagram, along with other evolving properties.[34] Accurate models can be
used to estimate the current age of a star by comparing its physical properties with those of stars along a
matching evolutionary track.

Das könnte Ihnen auch gefallen